Vous êtes sur la page 1sur 31

appliqué

sciences
Article
Analyse thermique d’un barrage en béton en tenant
compte de l’insolation, de l’ombrage, du niveau d’eau
et des débordements
1, 2 et Andrej Kryžanowski
Pavel Žvanut *, Goran Turk 2

1 Institut national slovène du bâtiment et du génie civil, 1000 Ljubljana, Slovénie


2 Faculté de génie civil et géodésique, Université de Ljubljana, 1000 Ljubljana, Slovénie; goran.turk@fgg.uni-
lj.si (G.T.); andrej.kryzanowski@fgg.uni-lj.si (A.K.)
* Correspondance : pavel.zvanut@zag.si

Application : Les distributions saisonnières de température dans les barrages en béton affectent
les charges thermiques, qui peuvent causer des contraintes élevées dans le béton et affecter
significativement l’occurrence et le comportement des fissures, ainsi que des déplacements (en
particulier dans les barrages en arc et en gravité).

Résumé: Cette étude présente une procédure de modélisation du processus de transfert de


chaleur dans un barrage en béton, en tenant compte des conditions limites temporelles sur les
côtés amont et aval du barrage (c.-à-d. le niveau d’eau du réservoir, les débordements, l’insolation
et l’ombrage) qui affectent les conditions de température du barrage. Le grand barrage de Moste
(Nord-Ouest de la Slovénie) a été analysé, où un système automatisé de mesure des
températures du béton et de l’eau et de surveillance des effets météorologiques a été installé. Des
analyses thermiques (1D et 2D) pour la conduction thermique non linéaire et non stationnaire par
les solides ont été effectuées à l’aide d’un programme basé sur la méthode des éléments finis
(FEM), TeEx, qui a été complété par deux programmes spécialement développés pour déterminer
les effets de la convection et de l’insolation, considérant aussi l’effet d’ombrage. Une période de
15 jours a été analysée, ainsi qu’une période d’un an. Il a été constaté que les résultats des
analyses effectuées étaient bien adaptés aux mesures expérimentales de la température du
Citation : Žvanut, P.; Turk, G.; béton. Les résultats ont montré qu’à l’intérieur du barrage, le gradient de température était le plus
Kryžanowski, A. Thermal Analysis of grand dans une zone très étroite le long de la surface en béton, mais la température ne s’est pas
a Concrete Dam Taking into Account
stabilisée moins qu’à une profondeur d’environ 6 m. Dans le cadre des analyses thermiques, des
Insolation, Shading, Water Level and
analyses d’incertitude des résultats des calculs ont également été effectuées.
Spillover. Appl. Sci. 2021, 11, 705.
https://doi.org/10.3390/app11020705
Mots clés: barrage en béton; transfert de chaleur; rayonnement solaire; ombrage; convection;
Date de réception :
analyse thermique; méthode des éléments finis; surveillance de la température; champ de
23 décembre 2020 température; incertitude
Date d’acceptation : 10 janvier 2021
Publication : 13 janvier 2021

Note de l’éditeur :L’IPMD demeure 1. Introduction


neutre en ce qui concerne les
Ces derniers temps, la surveillance sanitaire des barrages en béton est devenue un
réclamations juridictionnelles dans
sujet de grande importance et implique la surveillance du comportement statique et
les cartes publiées et les affiliations
dynamique des grands barrages [1]. L’un des paramètres les plus importants qui influencent
institutionnelles.
fortement le comportement statique des barrages en béton, en particulier les barrages en arc,
est la température. Pour cette raison, il est très important de connaître les distributions
saisonnières de la température dans les barrages en béton, car ils affectent les charges
Copyright :© 2021 par les auteurs. thermiques, qui peuvent causer des contraintes élevées dans les structures de barrages en
Titulaire de licence MDPI, Bâle, béton et affecter considérablement l’occurrence et le comportement des fissures, ainsi que
Suisse. Cet article est un article en les déplacements. Il faut tenir compte des effets des changements climatiques et du
libre accès distribué en vertu des réchauffement climatique, qui peuvent entraîner une augmentation générale de la
modalités de la licence Creative température de ces structures [2]. Dilger et al. [3] et Carslaw et Jaeger [4] ont effectué des
Commons Attribution (CC BY)
recherches fondamentales préliminaires dans le domaine de l’analyse thermique liée à la
(https://
conduction de la chaleur non linéaire et non stationnaire par un solide, en tenant compte de
creativecommons.org/licenses/by/
conditions limites convenables, tandis que quelques années plus tard, Leger et al. [5,6 ] a
4.0/).
présenté une méthodologie basée sur la méthode des éléments finis (FEM), qui pourrait être
utilisée pour déterminer la barrages de gravité en béton. En utilisant cette méthodologie, le comportement thermique
température saisonnière et des structures de barrages en béton (de l’arche à la gravité) a été beaucoup mieux analysé,
la répartition des tant en
contraintes dans les

Appl. Sci. 2021, 11, 705. https://doi.org/10.3390/app11020705 https://www.mdpi.com/journal/applsci


Appl. Sci. 2021, 11, 705 2 de 19

la phase de conception du barrage [7,8] ainsi que la phase ultérieure de surveillance du


barrage [9,10]. Elle a été discutée dans un nombre croissant d’articles traitant de :
l’analyse des contraintes des structures en béton, qui sont affectées par des charges
thermiques variables [11]; la détermination du champ de température périodique dans un
barrage en béton [12]; le calcul numérique de la température dans un béton massif [13];
analyse thermique et des contraintes du barrage en béton compacté à rouleaux [ 14]; et
l’effet des impacts environnementaux sur l’analyse des contraintes thermiques du
barrage en béton voûte [15]. D’autres auteurs ont discuté de l’élaboration de nouveaux
modèles numériques, comme le modèle pour l’analyse des barrages en béton en raison
des effets thermiques environnementaux [16], le modèle hydrostatique et le modèle de
déplacement de température dans le temps pour les barrages en béton [17]. et modèle
d’état hydrostatique saisonnier pour l’analyse des données de surveillance des barrages
en béton [18]. D’autres ont fait état de l’étude de sensibilité des champs thermiques dans
les structures massives [19] et de la protection thermique des barrages en béton dans
les régions nordiques [20].
Ces dernières années, de nombreuses recherches ont été menées sur la performance
thermique des barrages en béton, en tenant compte des effets de l’évolution des conditions
environnantes. Certains chercheurs se sont concentrés sur des études détaillées de
déplacements portant sur : l’analyse de barrages en béton sur la base de charges
hydrostatiques saisonnières [21]; déplacements thermiques saisonniers des barrages de
gravité, situés dans des régions froides [22]; la méthode de déformation des barrages par
action thermique [23]; et la corrélation entre la température de l’air et la variation quotidienne
de la réponse structurelle [24]. D’autres chercheurs ont discuté des fissures dans les
structures des barrages [25,26], tandis que d’autres ont analysé les contraintes induites par
la température dans les barrages en béton [27,28]. D’autres encore ont étudié les effets
d’ombrage [29,30] et les effets du rayonnement solaire [31,32], et certains rapports ont
discuté des champs de température dans le cas des barrages à voûte très haute [33-36].
D’autres chercheurs ont récemment comparé les données de surveillance sur le
terrain avec les résultats d’analyses numériques des distributions de température
également pour d’autres structures (Xia et al. [37] pour les ponts suspendus à longue
portée, Su et al. [38] pour les structures superhautes) et pour des matériaux spécifiques
(Abid et al. [39] pour les poutres en acier enrobées de béton).
Une recension des écrits sur la recherche sur les méthodes d’analyse des données
de surveillance des barrages, qui sont très importantes pour surveiller la sécurité des
barrages et où l’influence des variables environnementales (p. ex., niveau d’eau,
température) a été impliquée, a récemment été présentée par Li et al. [40].
Toutefois, dans les articles susmentionnés, il ne semble pas y avoir de base de
données exhaustive sur les résultats des mesures de température au cours des
décennies qui ont suivi la construction initiale du barrage, ou sur la modification des
champs thermiques dans le béton en raison de changements quotidiens et saisonniers
des conditions limites.
Dans cet article, les auteurs présentent une procédure relativement simple qui peut
être utilisée pour modéliser le processus de transfert de chaleur dans un barrage en
béton, en tenant compte des conditions limites variables dans le temps, tant du côté
amont d’un barrage, comme l’effet de l’oscillation du niveau d’eau du réservoir, ainsi que
sur le côté en aval, insolé, tels que les effets de l’insolation, ombrage et l’eau se
déversant sur la crête du déversoir. Cette procédure a été vérifiée au barrage de béton
le plus élevé en Slovénie.

2. Matériaux et méthodes
2.1. Analyse thermique d’un barrage en béton
Dans le cas d’un barrage en béton avec rayonnement solaire incident utilisant l’effet
d’ombrage, un niveau d’eau variant dans le temps du réservoir et des débordements
occasionnels, le processus de transfert de chaleur est illustré à la figure1.
• Neumann ou flux de chaleur imposé (le flux de chaleur de surface défini, c’est-à-
dire l’absorption du rayonnement solaire compte tenu de l’effet d’ombrage);
• Robin ou flux convectif (le flux de chaleur dépend linéairement de la différence de
température entre la surface et la température ambiante, par exemple, l’échange de chaleur par
Appl. Sci. 2021, 11, 705 convection d’air ou par convection d’eau); 3 de 19
• Rayonnement provenant de la surface d’un solide (le flux de chaleur dépend non
linéairement de la différence de température entre la surface et la température ambiante).

Figure 1.Processus de transfert de chaleur dans un barrage en béton.


Figure 1.Processus de transfert de chaleur dans un barrage en béton.
TheHeatcorrespondingexchangebyequationconvectionwhichcanrelatesbedeterminednonlinear andbythenonstatiowell-knownaryhequationconducpre--

tion sentedinthebycaseëvanutofaettwoal. -dimensional[41]:espace et un solide isotrope homogène dont


la conductivité thermique est indépendante de la température [4] est :
qc = hc (T T a)
T T
2 2 ρ c T
où qcis le flux convectif (m/h+ 2); h c =est la co nv le co
x2 y2 λ t

température de surface du barrage (K); et Tais la température ambiante (K).


L’effet des radiations a été vérifié et jugé faible, de sorte qu’il a été négligé. Où T est la température (K); x, y sont les
coordonnées cartésiennes (m); ρest la densité
same 3a été constaté concernant l’effet du refroidissement du béton dû à l’évaporation de (kg/m ); c est la chaleur spécifique (J/(kg K)); λest la conductivité thermique
(W/(m K)); et est
le(s) temps absorbé(s). arroser.
Les principaux types de descriptiondesprocédures de base sont proposés par Dilger et al. [3] pour la
• détermination assez précise de la densité du flux d’énergie due à l’absorption des rayons solaires
Dirichlet ou la température imposée (la température de surface définie, par exemple de la
Žvanut et al. [41], où la quantité d’énergie solaire absorbée par la fondation);
corps d’un système est donné par
• Neumann ou flux de chaleur imposé (le flux de chaleur de surface défini, c’est-à-dire l’absorption de
rayonnement solaire considérant l’effet de
=shading);Icosθ question
s

• Robin ou flux de convection (le flux de chaleur dépend linéairement de la différence de


température entre la surface et la température ambiante, par exemple, l’échange de chaleur par convection d’air ou par
convection d’eau);
• Rayonnement provenant de la surface d’un solide (le flux de chaleur dépend non
linéairement de la différence de température entre la surface et la température ambiante).
L’échange de chaleur par convection peut être déterminé par l’équation bien
connue présentée par Žvanut et al. [41] :
qc = hc (T Ta)
); h
où qcis le flux convectif (W/m2 cis le coefficient de convection (W/m2K); T est la température de
la température ambiante (K).
surface du barrage (K); et Tais
L’effet des radiations a été vérifié et jugé faible, de sorte qu’il a été négligé. Il en a
été de même pour l’effet de refroidissement du béton dû à l’évaporation de l’eau
absorbée.
Une description d’une procédure relativement simple proposée par Dilger et al. [3]
pour la détermination assez précise de la densité de flux d’énergie due à l’absorption du
rayonnement solaire est présentée par Žvanut et al. [41], lorsque la quantité d’énergie
solaire absorbée par le corps d’un système est donnée par
qs = I · a · cos θ
Appl. Sci. 2021, 11, 705 4 de 19

); I est la
où qsis la densité de flux d’énergie du rayonnement solaire absorbé par le solide (W/m2
densité de flux d’énergie du rayonnement solaire à la surface de la Terre (W/m2); a est l’absorption
solaire de la surface; θ
est l’angle d’incidence des rayons du soleil ( );

I = Isc · kT
); et kT
où Iscis la constante solaire (en moyenne 1350 W/m2 is le facteur de transparence
(dépend des conditions atmosphériques et de la longueur du trajet du rayonnement
dans l’atmosphère);
T

(k m t U )
k = 0,9 A

où kAis facteur d’altitude H (où H est mesuré en mètres); m est le facteur d’influence de la longueur
relative du trajet du rayonnement (selon l’angle d’élévation solaire βs); et tUest le facteur de pollution
atmosphérique (entre 1,8 pour l’air pur et 9,0 pour l’air très pollué);
kA − H
= 1 0.000105
1
m= , si βs 5◦
sinβs ≥

1

m= , si βs <5
sin5
L’angle d’incidence des rayons du Soleil peut être présenté par

cosθ = −cosδ · cosΩ · sinα · cosτ · sinΦ + cosδ · sinΩ · sinα · sinτ

+sinδ cosΩ sinα cosΦ+cosδ cosα cosτ cosΦ+sinδ cosα sinΦ

et cosθ 0
où δest la déclinaison (c’est-à-dire l’angle entre le plan équatorial et la direction du Soleil;
positif vers le pôle céleste nord); Ω est l’azimut de la normale au plan (mesuré dans le
sens horaire à partir du nord); αest l’angle entre le plan et la surface de la Terre; Φest la
latitude géographique; et τest l’angle horaire (positif le matin), qui est indiqué par

τ= (12 u) 15
par heure).
où u est l’heure de la journée (en 24 heures; l’angle horaire change à raison de 15
Dans le cas du plan vertical (α = 90), l’équation (8) peut être simplifiée en
cosθ = −cosδ · cosΩ · cosτ · sinΦ + cosδ · sinΩ · sinτ + sinδ · cosΩ · cosΦ (10)

), il peut être simplifié en


que, dans le cas du plan horizontal ( α= 0

cosθ = cosδ · cosτ · cosΦ + sinδ · sinΦ = sinβs


où βsis l’angle d’élévation solaire.
L’angle d’azimut solaire (αs) définit la direction du Soleil, en supposant la
convention dans le sens nord-horaire, et il peut être défini par l’expression

cosαs = sinδ−sinβs ·sinΦ


cosβs ·cosΦ

si τ > 0 (u < 12, le matin) αs < 180

si τ < 0 (u > 12, dans l’après-midi) αs> 180


L’équation de conduction thermique (1), avec des conditions limites bien définies,
peut être résolue numériquement par le FEM [42]. Une telle discrétisation du modèle en
fini
Appl. Sci. 2021, 11, 705 5 de 19

éléments résulte en un système d’équations différentielles non linéaires de premier ordre.


Dans le cas de la recherche de températures de noeud dépendant du temps, le système
prend la forme suivante :

K T + C T, t = F
où Kis la matrice de conductivité thermique; Test du vecteur de température nodale; Cis
la matrice de capacité de chaleur; T,t est la dérivée temporelle de la température; et Fis
le vecteur des actions extérieures.
Le côté gauche de l’équation (13) et les coefficients dépendent, en général, des
températures recherchées (T), tandis que le côté droit est généralement aussi une
fonction explicite du temps (t).

K (T) T + C (T) T, t = F (T, t)


2.2. Travail sur le terrain
Le grand barrage de la Moste, construit en 1952 sur la rivière Sava Dolinka, au
nord-ouest de la Slovénie, a été analysé. C’est le plus haut barrage de Slovénie, avec
une hauteur structurale de 59,80 m. La crête du barrage a une longueur de 72,00 m,
tandis que le volume du barrage est de 42000 m3. La crête du déversoir est à une
altitude de 523,50 m. Cependant, les vannes en acier, qui ont une hauteur de 1,25 m,
augmentent la crête effective du déversoir à l’altitude de 524,75 m. Le barrage est situé
, alors que
à la latitude nord de 46,41 l’azimut de l’axe symétrique du côté aval du barrage
, de sorte qu’il fait pratiquement face au sud.
de la Moste est de 186
Un système automatisé de mesure des températures du béton et de l’eau et de
surveillance des effets météorologiques à l’aide d’une station météorologique
automatique mobile (MAWS) a été installé sur le barrage en juillet 2013. Un système de
mesure automatique du niveau d’eau du réservoir et de la hauteur de l’eau déversée sur
la crête du déversoir avait été établi en 2000.
Avant de déterminer l’emplacement des jauges de température du béton et des
jauges de température de l’eau, des calculs préliminaires et des analyses des résultats
de la conduction de la chaleur dans la structure du béton ont été effectués, ainsi qu’une
analyse préliminaire des données disponibles concernant l’oscillation du niveau d’eau
du réservoir de la Moste au cours des 30 dernières années.
Three boreholes were drilled in the dam and six concrete temperature gauges were
installed. In the upper borehole on the downstream side of the dam, which was made where

the dam surface inclination is 30 , four gauges (TC1 to TC4) were installed at different
distances from the dam surface: TC1 at 0.1 m (520.3 m a.s.l.), TC2 at 0.2 m (520.2 m a.s.l.),
TC3 at 0.5 m (520.0 m a.s.l.) and TC4 at 1.0 m (519.6 m a.s.l.). In the lower borehole on the

downstream side of the dam, which was made where the dam surface inclination is 43 , one
gauge (TC6) was placed at a distance of 0.1 m from the dam surface (485.2 m a.s.l.). An
additional concrete temperature gauge (TC5) was installed 0.3 m from the dam surface
(521.2 m a.s.l.) in a borehole on the upstream side of the dam.
The water temperature measurements were performed using four gauges (TW1 to
TW4), placed in a reservoir in a steel pipe, which protected them against floating debris
and rushing waters, located at altitudes 519 m, 521 m, 522 m and 523 m. The lowest
gauge was installed at a depth that was not affected by the surface zone of changing
water temperatures [43], which was confirmed by an additional gauge placed in the
reservoir at altitude 511 m.
The measurements of meteorological effects included the following parameters:
solar radiation, precipitation, air temperature, wind speed and wind direction. The
location of the MAWS, at an altitude 527.65 m, was determined after a detailed visual
inspection of the area near the dam, bearing in mind easy accessibility, as few
topographical obstacles as possible and low noticeability.
An overview of the locations of the concrete and water temperature gauges as well
as the MAWS is given in Figures 2–4.
Appl. Sci. 2021, 11, x FOR PEER REVIEW 6 of 20

of the area near the dam, bearing in mind easy accessibility, as few topographical obsta-of the area near the
dam, bearing in mind easy accessibility, as few topographical obstacles as possible and low noticeability.
Appl. Sci. 2021, 11, 705 6 of 19
cles as possible and low noticeability.
An overview of the locations of the concrete and water temperature gauges as well
An overview of the locations of the concrete and water temperature gauges as well
as the MAWS is given in Figures 2–4.
as the MAWS is given in Figures 2–4.

Figure 2. The temperature monitoring system established in the central cross-section of the Moste
Figure 2. The temperature monitoring system established in the central cross-section of the Moste Dam.

Figure 2. The temperature monitoring system established in the central cross-section of the Moste Dam.
Dam.

Figure 3. The locations of the mobile automatic weather station (MAWS) and of the temperature
Figure 3. The locations of the
mobile automatic weather station (MAWS)
and of the
temperature

Figure 3. The locations of the and of the gauges on the downstream side of the dam
(Photo: P. Žvanut).
gaugesononthethedownstreamsidesideofofthethedamdam(Photo:.PŽvanut)...
Appl. Sci. 2021, 11, 705 7 of 19
Appl. Sci. 2021, 11, x FOR PEER REVIEW 7 of

20

Figure 4. The locations of the temperature gauges on the upstream side of the dam and in the pro-
Figure 4. The locations of the temperature gauges on the upstream side of the dam and in the
tective pipe (Photo: P. Žvanut).
protective pipe (Photo: P. Žvanut).

2.23.. 3Procédure.pourlesAnalyses Thermales


2.3.1. General
2.3.1. General
The temperature conditions of the dam were determined by means of the procedure The temperature conditions of the
dam were determined by means of the procedure
developed for modeling the heat transfer process in concrete dams, taking into account developed for modeling the heat transfer process in concrete
dams, taking into account
time-varying boundary conditions on both the upstream side of the dam (the effect of time-varying boundary conditions on both the upstream side of
the dam (the effect of the
the oscillation of the water level of the reservoir) and on the downstream side (the effects oscillation of the water level of the reservoir) and on the
downstream side (the effects of
of insolation, shading, and water spilling over the spillway crest). Computer programs insolation, shading, and water spilling over the spillway crest).
Computer programs for
for the determination of temperature during time, at each node of the finite element the determination of temperature during time, at each node of the
finite element mesh,
mesh, were developed within the programming environments Mathematica [44] and were developed within the programming environments
Mathematica [44] and MATLAB
MATLAB [45]. The thermal properties of the mass concrete used in the determining of the [45]. The thermal properties of the mass concrete used in
the determining of the temper-
temperature conditions of the dam are given in Table 1.
ature conditions of the dam are given in Table 1.
Table 1. The thermal properties of the mass concrete.
Table 1. The thermal properties of the mass concrete.
1
Material Property Unit Expected Value Standard Deviation 1
Material Property Unit Expected Value Standard Deviation

3
Density Density kg/m3 kg/m2400 2400 96 96
Specific heat J/(kg K) 960 96
Specific heat J/(kg K) 960
96
Thermal conductivity W/(m K) 2.33 0.233
Thermal conductivity W/(m K) 2.33 0.233
2
Convection coefficient (air) W/(m K) 55.6 11.12
Convection coefficient (air) W/(m 2K) 2 55.6 11.12
Convection coefficient (water) (m) K) 556 55.6
2
Convection coefficient (water)Solar absorptivityW/(m K) / 556 0.65 55.60.0325
/
Solar absorptivity Emissivity / / 0.65 0.90 0.0325
Emissivity / 0.90 /

1 Taken into account in determining the uncertainty of the calculation results.


1 Taken into account in determining the uncertainty of the calculation results.
2.3.2. Effect of Spillover
2.3.2 .Effect of Spillover
Spilling of waterfrom the reservoir over the spillway crest has an effect on the
determination of the ofconvection (i .e., thetemperature the
Spilling of water parametersfromthe reservoir over the spillway cresthasand effect convectionthede-
coefficient)terminationandof ofthetheparametersheatfluxdueofconvectiontoinsolation,(i.etaking.,thetemperatureintoaccountandthe

theeffeconvectionofshading,co-efficient) and of the heat flux due to insolation, taking into account the effect of shading,
certain time
due to spillover, the water temperature and water convection coefficient are on selected surface, at a given time. In the case that the selected surface is
wetted at a
are
considered; however, while there is no heat flux due to the insolation. Otherwise, if at certain time due to spillover, the water temperature and water convection
coefficient
certain time the selected surface is dry (i.e., there is no spillover), the parameters of air considered; however, while there is no heat flux due to the
insolation. Otherwise, if at a
convection (i.e., the air temperature and the air convection coefficient) and the measured certain time the selected surface is dry (i.e., there is no
spillover), the parameters of air
heat flux due to insolation are taken into account.

convection (i.e., the air temperature and the air convection coefficient) and the measured
During the analyzed year, 1306 h of water spilling over the spillway crest were recorded heat flux due to insolation are taken into account.
(from November to May), which is equivalent to 14.9% of the year. The longest period During the analyzed year, 1306 h of water spilling over the
spillway crest were rec-
of continuous spillover was recorded between 14 March 2014 at 23:00 and 25 March 2014 orded (from November to May),
which is equivalent to 14.9% of the year. The longest
period of continuous spillover was recorded between 14 March 2014 at 23:00 and 25
Appl. Sci. 2021, 11, 705 8 of 19
Appl. Sci. 2021, 11, x FOR PEER REVIEW 8 of 20

atMarch7:00,and2014amountedat7:00,andto amounted249h,whereasto249theh, longestwhereasperiodthelongestofcontinuousperiodof spillovercontinuousat a

spilloverheight ataspilloverof 5heightcmwasof recordedover5cmbetweenwasrecorded12Maybetween2014at 1214:00Mayand201421 atMay14:002014

atand15:00,21 Mayand amounted2014at15:00,to 218and hamounted(Figure5to). 218 h (Figure 5).

526 Water level of the reservoir Spilling of water

525 524.75

524
523.50
l. )

523
Altitude (m a. s.

522

521

520

519

518
1.9.13 31.10.13 31.12.13 2.3.14 2.5.14 2
Date

Figure5.5The.waterlevelof the reservoir and the correspondingheightofofthethespillover..


2.3.3. Effect of Shading
2.3.3. Effect of Shading
In order to investigate the effect of shadingg (i.e., solar exposure or non-exposure of the
In order to investigate the effect of (i.e., solar exposure or non-exposure of
observation point on the downstream insolated side of the dam), a method developed by the observation point on the downstream insolated side of the dam), a method

ŽvanutbyŽvanutetal.et[41al,.46[41,46]—based—basedonterrestrialonterrestriallaserlaserscannerscannermeasurementsof theofthetopography-ofraphythewideroftheareawiderof

thareadam,oftheanddam,of andthe useoftheof usetwoofcomputertwocomputerprogramsprograms(developed(developedwithin

thewithincomputingthecomputingenvironmentsenvironmentsMathematicaMathematicaandMATLAB,andMATLAB,respectively)respectively)—was—appliedwas.

Usingappliedthis.Usingapproachthis approachitispossibleitis possible tdetermintodetermine—forany—forselectedanyselectedobservationobservationpoit—


thepointelevation—theelevationanglesofanglestheterrainofthe atterraindifferentatdifferentazimuthsazimuths(i.e.,the(i .econtour.,thecontourofthe ofterrain),the

theterrain),positiontheofpositiontheSun,ofandthe insolationSun,andinsolationovertimeover.Thetimepropos.Thedproposedmethod hasmethodbeenhasconfirmedbeen

asconfirmedhighlyreliableashighlyinthreliablecase ofin the casedownstreamofthedownstreamsideofthesideMosteoftheDamMoste[46]Dam. [46].


6 as from the point P1, Figure6 shows the contour of the terrain asseen from theobservation point P1,

which is side of the dam (Figure 3) at altitude which is locatednear the top of the downstream side of the dam (Figure 3) at altitude

Appl. Sci. 2021, 11, x FOR PEER REVIEW 9 of 20


520.35 m, and of the Sun in both extreme positions during the year (i. ., at

520.35 m, and the virtual path of the Sun in both extreme positions during the year (i.e.,
atthethesummersummerandandwinterwintersolstices)solstices)..

70

60
Terrain contour from P1
50
Path of the Sun on June 21

40 Path of the Sun on December 21

30

20

10

0
8 10 12 14 16 20 22 24 26 Azimuth (°)
0 0 0 0 0 180 0 0 0 0
6. as from po P1 and the virtual path of the Sun .
Figure6 . The contour of the terrain asseen from point P1 and the virtual path ofthe Sun .

2.3.4. Solar Radiation


Back-analyses of the solar radiation were performed for the analyzed period of 15
consecutive clear days in the summer at the location of the MAWS, where the calibrated
parameters were obtained [41].
Appl. Sci. 2021, 11, 705 9 of 19

2.3.4. Solar Radiation


Back-analyses of the solar radiation were performed for the analyzed period of 15
consecutive clear days in the summer at the location of the MAWS, where the calibrated
parameters were obtained [41].
Figure 7 shows the calculated absorbed heat flux due to solar radiation at observation
point P1 on the downstream side of the dam in the case of year-round insolation and the
measured absorbed heat flux obtained by the measurements of the insolation at the MAWS.
The calculated values, which represent the theoretical daily maximums of the heat flux,
consistently follow the measured values in the analyzed year, but due to different effects (i.e.,
cloudiness, partial transparency of the atmosphere) are higher. A greater exception occurs in
the winter time, when the measured values are significantly lower than the calculated values
in the case of year-round insolation. The main reason for this deviation concerns the
difficulties in the insolation measurements at low positions of the Sun (small elevation angles
of the Sun and therefore greater lengths of the path of the radiation through the atmosphere)
in the winter. The maximum daily maximum (annual maximum) of calculated absorbed heat
2
flux in the case of year-round insolation during the analyzed year was 630 W/m in the
summer period, whereas the minimum daily maximum of calculated absorbed heat flux was
2
284 W/m in the winter period (the blue spikes show the daily
Appl. Sci. 2021, 11, x FOR PEER REVIEW 10 of 20 maximums during the year). Figure 7 also shows the zero-values of the heat fluxes of
solar
radiation during the actual water spilling over the spillway crest (the gaps in the timeline).

700
P1 - Calculated (Sun all year round)
600 P1 - Measured (MAWS)

500

400

300

200

100

0
1.9.13 31.10.13 31.12.13 2.3.14 2.5.14 2.7.14 1.9.14
Date

Figure 7. The calculated absorbed heat flux due to theoretical year-round insolation and the measured Figure 7. The
calculated absorbed heat flux due to theoretical year-round insolation and the
absorbed heat flux at point P1.

measured absorbed heat flux at point P1.


3. Results and Discussion
3. Results and Discussion
3.1. General
3.1. General
The thermal analyses (1D and 2D) were carried out by using the FEM-based computer
The thermal analyses (1D and 2D) were carried out by using the FEM-based com- program TeEx (written in
MATLAB, [47]) which was complemented by two specially puter program TeEx (written in MATLAB, [47]) which was
complemented by two spe-formulated programs for determining the effects of convection and insolation, taking into
cially formulated programs for determining the effects of convection and insolation, account the effect of shading.
taking into account the effect of shading.
Since thesame initial concrete
Since the same initial concrete temperatures were not the proper approximation of actual conditions atthe beginning of
temperatures werenot theproper approximation of
analyses, thefollowing procedure was implemented .
Theactualinitialconditionsconcreteattemperaturesthebeginningwereofanalyses,obtainedthein followingsuchawayprocedurethatusingwaspreliminaryimplement-


initialed.Thetmperaturesinitialconcreteof20temperatures C,thetemperatureswereobtainedafterinfivesuchdaysway(forthattheusing1Danalyses)preliminaryo

afterinitialonetemperaturesyear(forthe of2D20analyses)°C,the weretemperaturesfirstcalculated,afterfiveanddaysthen(forurtherpetitions1Danalyses)ofthisor

periodafteronewereyearper(formed,the2Dwherebyanalyses)thewererespectivefirstcalculated,tempraturesandthenatthefourendrepetitionsofeachperiodofthis

wereperiodusedwereastheperformed,initialvalueswherebyofthe thenextrespectiverepetitiontemperatures.Inordertoeliminatetheendthe ofimpacteach ofperiodthe

were used as the initial values of the next repetition. In order to eliminate the impact of the preliminarily selected initial temperatures, the initial values of the

actual calculation were therefore defined as the temperatures calculated at the end of the final repetition.
The 15-days period (Section 3.2.1) and the one year period (Section 3.2.2) were first
analyzed using a 1D model, represented by a line perpendicular to the concrete surface,
for the instrumented locations of the Moste Dam (Figure 2). In both periods the change-
Appl. Sci. 2021, 11, x FOR PEER REVIEW 11 of 20

Appl. Sci. 2021, 11, 705 10 of 19

(TC5), considering the water level of the reservoir, and also the measured air temperatures
(or water temperatures, Figure 10) and the height of spillover during the analyzed
preliminarily selected initial temperatures, the initial values of the actual calculation were 15-days period, from 10 May to 25 May 2014.
therefore defined as the temperatures calculated at the end of the final repetition.
It can be seen from Figures 8 and 9 that the calculated concrete temperatures at the The 15-days period (Section 3.2.1) and
the one year period (Section 3.2.2) were first
downstream gauges TC1 and TC4, when taking into account the effect of spillover, are in analyzed using a 1D model,
represented by a line perpendicular to the concrete surface,
very good agreement with the measured concrete temperatures during the whole period for the instrumented locations of the
Moste Dam (Figure 2). In both periods the changeable
(the root mean square error (RMSE) is 1.95 at TC1 and 1.73 at TC4, whereas the mean cloudiness with the effect of shading of the
dam surface, and the effect of the varying
absolute error (MAE) is 1.39 at TC1 and 1.67 at TC4).
water level of the reservoir with the occasional spilling of water over the spillway crest The measured and calculated concrete temperatures at the
upstream gauge TC5
were taken into account [48]. The heat flux of the absorbed solar radiation was determined (Figure 10), during the period of spillover, or more
precisely, during the period when the
from the results of the MAWS measurements. Secondly, the temperature fields of the
water level of the reservoir is above the altitude of gauge TC5, are in very good agree- Moste Dam, during the year, were
determined using a 2D model of the symmetrical cross-
ment (RMSE is 0.45 and MAE is 0.31).
section of the complete dam, where all the previously mentioned effects were taken into
From Figure 11 it can also be seen that there is very good agreement between the account (Section 3.3). Thirdly, uncertainty
analysis of the calculated temperature field of
measured and calculated concrete temperatures at gauge TC6, taking into account the the Moste Dam on 1 August 2014 at 12:00
was performed, using normally distributed
effect of spillover, assuming that the height of the spillover must be at least 0.1 m, oth- random variables (Section 3.4).
erwise the exposed surface of the lower downstream borehole is not wetted (RMSE is 1.28
3.2and.EffectMAEof Spilloveris1.07).
3.2.1. 15Figure-Days 3Periodshows the minor amount of spillover (when the height of spillover is less
than 0.1 m), where it can be seen that the exposed surface of the upper borehole, with Figures 8–11 show the measured
and calculated concrete temperatures at three down-
gauges TC1 to TC4, is wetted irrespective of the height of spillover, whereas the exposed stream gauges (TC1 and TC4 located in
the upper borehole, and TC6 located in the lower
surface of the lower borehole, with gauge TC6, is not wetted at a spillover height of less borehole), with and without the effect of
spillover, and at one upstream gauge (TC5),
than 0.1 m, since it is located outside of the central cross-section of the dam and therefore considering the water level of the
reservoir, and also the measured air temperatures (or
not reached by the wetted area of the downstream side of the dam.

water temperatures, Figure 10) and the height of spillover during the analyzed 15-days
The maximum (or minimum) temperature at the specific point in concrete occurs period, from 10 May to 25 May 2014.
later than the maximum (or minimum) air temperature. This time difference is termed It can be seen from Figures 8 and 9 that the
calculated concrete temperatures at the
the time lag and depends on the distance from the concrete surface. At a depth of 0.1 m downstream gauges TC1 and TC4, when taking into account the effect of
spillover, are in
(gauges TC1 and TC6) the time lag was fairly short (at maximum temperatures 1 to 2 h, very good agreement with the measured concrete
temperatures during the whole period
and atminimum 3to 4 h) . At depth of0 . 2 m (gauge TC2) thetime lagwas
(the root meansquaretemperatureserror(RMSE) is 1 .95 at TC1and 1 .73 atTC4, whereas the mean

absoluteslightlyerrorlonger(MAE)(atmaximumis1.39at TC1temperaturesand1.67at2 TC4)to3h,. and at minimum temperatures 4 to 5 h). TheAta

measureddepthof0.and3m (gaugecalculatedTC5)concretethetimetemperatureslagthemaximumatthe upstreamtemperaturesgaugewasTC5to 6

(Figureh,and106),toduring8hat the periodminimumofspillover,temperaturesormore.Atprecisely,adepth

ofduring0.5mthe(gaugeperiodTC3)whenthethetime waterlag waslevelalreadyofthereservoir10to12 ish,aboveandatthea

altitudedepthofof1 .gauge0m(gaugeTC5,areTC4)in verythe timegoodlagagreementwaseven

(RMSEseveralis days0.45and. MAE is 0.31).

40
72
35
TC1 - Measured
30 64
TC1 - Calculated (no spillover)
25
TC1 - Calculated (spillover) 56
20
(cm)

Air (MAWS) 48
15 Spilling of water
spillover

10
40
5
32
0
of
Heigh

-5
t

24
10.5.14
16

8
0 13.5.14 16.5.14 19.5.14 22.5.14 25.5.14 Date

FigureFigure8. 8.TheTheconcreteconcretetemperaturestemperaturesat atgaugegaugeTC1TC1duringduringthetheanalyzedanalyzed15-15day-dayperiod.period.


Appl. Sci. 2021, 11, 705 11 of 19
Appl. Sci. 2021 , 11 , x FOR PEER REVIEW 12 of 20
Appl. .Sci. 2021, 1111, xxFORPEERREVIEW 1212de2020

40
40 72
35 TC4 - Measured 72
35
TC4 – Mesuré
30
30 64
25 TC4 - Calculated (no spillover) 64
25
TC4--Calculé (nospillover)
20
20 TC4 - Calculated (spillover) 56
15 TC4--Calculé (débordement) 56

(cm)(cm)
15
48
10
10 Air (MAWS)
5 Air(MAWS) 48
55
Spilling of water

ofspilloverofof
0
00 Spillingofofwater 40

---555
10.10.5.5.14.5.14 40
32

HeightHeight
32
24
24
16
16
8
88
0
00

13.5.14 16.5.14 19.5.14 22.5.14 25.5.14


13.5.5.14. 1616.5.5.14. 19.5.5.14. 2222.5.5.14. 25.5.5.14.
Date

Date
Figure 9. The concrete temperatures at gauge TC4 during the analyzed 15-day period.
Figure99.9.The.concretetemperaturesatatatgaugeTC44during4thetheyzed1515-day--dayperiod...

40
40 TC5 - Measured 528
35 TC5 – Mesuré 528
TC5 - Calculated
35 TC5 – Calculé 527
30 TW2 - Measured (air/water)

30 TW2 – Mesuré (air/eau) 527


Water level of the reservoir
25 Waterlevelofofthereservoir 526
25
526
20
20 l. )l. )l. )

15 525
15
.

525
10
s.
a. s.a.a.s

10
5 524
55 TW2
(m(m(m)

0 TW2 524
00 523
AltitudeAltitude

---555
13.5.14 16.5.14 19.5.14
522

10.10.5.5.14.5.14 523
13.5.5.14. 16.5.5.14. 1919.5.5.14.
Date
522
Date
521
521
520
520
519
519

22.5.14 25.5.14
22.5.5.14. 25.5.5.14.

FigureFigure101010... L’opération de mesure des températures concrètes TC5TC555 pendant la période d’analyse de 515 jours--période de jour...
40
40 72
35 TC6 - Measured 72
35 TC6 – Mesuré
30
30 64
25 TC6 - Calculated (no spillover) 64
25 TC6--Calculé (nospillover)
20
TC6 - Calculated (spillover) 56
20
15 TC6--Calculé (débordement) 56

(cm)(cm)
15
10
TC6--Calculé (débordement>>10 cm) 48

TC6 - Calculated (spillover > 10 cm)


10
5 48

spilloverspillover
55 Air temperature (upstream)
0
00
Airtemperature(upstream)
---555 40
10.10.5.5.14.5.1
4 Spilling of water 40

Spillingofofwater
32

ofo
fof
32

Heig
htHe
ight
24
24

16

Height of spillover = 10 cm 16
Augmentation du débordement==10cm

8
88
0
00

13.5.14 16.5.14 19.5.14 22.5.14 25.5.14


13.5.5.14. 16.5.5.14. 19.5.5.14. 22.5.5.14. 25.5.5.14.
Date

Date
Figure 11The. concretetemperatures at gauge TC6 during theanalyzed 15-dayperiod.
Figure1111... concrete atatatgaugeTC666lors de la 1515-jour-jour ...
Appl. Sci. 2021, 11, 705 12 of 19

From Figure 11 it can also be seen that there is very good agreement between the measured
and calculated concrete temperatures at gauge TC6, taking into account the effect of spillover,
assuming that the height of the spillover must be at least 0.1 m, otherwise the exposed surface of
the lower downstream borehole is not wetted (RMSE is 1.28 and MAE is 1.07).
Figure 3 shows the minor amount of spillover (when the height of spillover is less
than 0.1 m), where it can be seen that the exposed surface of the upper borehole, with
gauges TC1 to TC4, is wetted irrespective of the height of spillover, whereas the
exposed surface of the lower borehole, with gauge TC6, is not wetted at a spillover
height of less than 0.1 m, since it is located outside of the central cross-section of the
dam and therefore not reached by the wetted area of the downstream side of the dam.
The maximum (or minimum) temperature at the specific point in concrete occurs later
than the maximum (or minimum) air temperature. This time difference is termed the time lag
and depends on the distance from the concrete surface. At a depth of 0.1 m (gauges TC1
and TC6) the time lag was fairly short (at maximum temperatures 1 to 2 h, and at minimum
temperatures 3 to 4 h). At a depth of 0.2 m (gauge TC2) the time lag was slightly longer (at
maximum temperatures 2 to 3 h, and at minimum temperatures 4 to 5 h). At a depth of 0.3 m
(gauge TC5) the time lag at the maximum temperatures was 5 to 6 h, and 6 to 8 h at the
minimum temperatures. At a depth of 0.5 m (gauge TC3) the time lag was already 10 to 12 h,
and at a depth of 1.0 m (gauge TC4) the time lag was even several days.
Appl. Sci. 2021, 11, x FOR PEER REVIEW 13 of 20
3.2.2. One Year Period
Figures 12 and 13 show the measured and calculated concrete temperatures at two of
the downstream gauges (TC3 and TC4, located in the upper borehole), with and without 3.2.2. One Year Period
the effect of water spilling over the spillway crest, and the height of spillover during the Figures 12 and 13 show the measured and calculated concrete
temperatures at two
analyzed one year period, from 1 September 2013 to 1 September 2014. It can be seen from of the downstream gauges (TC3
and TC4, located in the upper borehole), with and these figures that the calculated concrete temperatures at the downstream gauges
TC3
without the effect of water spilling over the spillway crest, and the height of spillover and TC4, when considering the
effect of spillover, are in very good agreement with the during the analyzed one year period, from 1 September 2013 to 1
September 2014. It can measured concrete temperatures during the whole year (RMSE is 2.50 at TC3 and 2.00 at be seen
from these figures that the calculated concrete temperatures at the downstream TC4, whereas MAE is 2.20 at TC3 and
1.84 at TC4).
gauges TC3 and TC4, when considering the effect of spillover, are in very good agree- The annual concrete temperature
oscillations for the analyzed year, at gauges TC1
ment with the measured concrete temperatures during the whole year (RMSE is 2.50 at to TC6, as well as the corresponding air
(or water) temperature oscillations are shown in
TC3 and 2.00 at TC4, whereas MAE is 2.20 at TC3 and 1.84 at TC4).

Tables 2–4.

40

35 72
TC3 - Measured
30 TC3 - Calculated (no spillover) 64
25 TC3 - Calculated (spillover) 56
(cm)

48
20
Spilling of water
15
spillover

10

5 40

0 32
of

-5
Heig
ht

1.9.13
24

16

0
31.10.13 31.12.13 2.3.14 2.5.14 2.7.14 1.9.14
Date

Figure1212.. La tempéra-t ure concrète escomptée pour la période de l’année analysée..


40
TC4 -
35 Measured
TC4 -
30 Calculated
(no
25 spillover)

20 TC4 -
Calculated
15 (spillover)
Spilling of
water
72

64

56

(cm)
48

spillover
40
32
H
5
0

-5
1.9.13 31.10.13 31.12.13 2.3.14 13 of 19
Appl. Sci. 2021, 11, 705
Date
Figure 12. The concrete temperatures at gauge TC3 during the analyzed year.

40 72
TC4 - Measured
35 TC4 - Calculated (no spillover) 64
30 TC4 - Calculated (spillover) 56

(cm)
Spilling of water

25 48

spillover
20 40

15 32

ht of
Heig
10 24

5 16

0 8

-5 0
1.9.13 31.10.13 31.12.13 2.3.14 2.5.14 2.7.14 1.9.14
Date

Figure 13. The concretetemperaturesatatgauggaugeTC4TC4duringduringthetheanalyzedyearyear..

The annual concrete temperature oscillations for the analyzed year, at gauges TC1 to
Table 2. The temperature oscillations and time lags at gauges TC1 to TC4.
TC6, as well as the corresponding air (or water) temperature oscillations are shown in

Tables 2–4.

Temperature ( C)
Limit The annual air temperature oscillations at the location of the MAWS were, for the
Air TC1 TC2 TC3 TC4
selected period, between −9.5 and 35.6 °C, so their amplitude was 45.1 °C. At gauge TC1,

1 1 2 1 2 1 2 1 2
theManalyzed Mannual oscillationsC wereM significantlyC smallerM(measuredC 35.3 °CMand calcuC-
lated 35.9 °C), whereas at gauge TC2 the oscillations were 30.7 and 29.4 °C, respectively.

Annual T min −9.5 0.9 −4.6 1.4 −2.4 2.9 1.1 4.9 3.5
T max 35.6 36.2 31.3 32.1 27.0 26.9 21.7 23.3 20.4
oscillations
Diff. 45.1 35.3 35.9 30.7 29.4 24.0 20.6 18.4 16.9

3 3 3
T min / 3–4 4–5 10–12 Several days
Time lag
3 3 3
T max / 1–2 2–3 10–12 Several days
1 M: measured.
2
C: calculated.
3
Time lag in hours.

Table 3. The temperature oscillations and time lags at gauge TC5.



Temperature ( C)
1 1 2
Limit Air Water Water
3 3 3 3
M M M M
Annual T min −1.8 2.7 2.9 3.8
T max 32.1 19.0 16.6 17.6
oscillations
Diff. 33.9 16.3 13.7 13.8
T min / / /
Time lag
T max / / /
1 À la jauge TW2 (c.-à-d. 521 m a. s. l.).
2 3
Jauge TW1 (c.-à-d. 519 m a. s. l.). M: measured.
4
C: calculated.

Table 4. The temperature oscillations and time lags at gauge TC6.



Temperature ( C)
Limit Air TC6

1 1
M M
T min −5.8 1.0
Annual oscillations T max 30.1 25.7
Diff. 35.9 24.7

T min / 3à4h
Time lag
T max / 1à2h
1 M: measured.
2
C: calculated.
Appl. Sci. 2021, 11, 705 14 of 19

The annual air temperature oscillations at the location of the MAWS were, for the
◦ ◦
selected period, between −9.5 and 35.6 C, so their amplitude was 45.1 C. At gauge TC1,

the analyzed annual oscillations were significantly smaller (measured 35.3 C and calculated
◦ ◦
35.9 C), whereas at gauge TC2 the oscillations were 30.7 and 29.4 C, respectively. At

gauge TC3 the measured annual oscillations were still 24.0 C, and the calculated oscillations

up to 20.6 C. At gauge TC4 the measured oscillations, over the analyzed annual period,
◦ ◦
were 18.4 C, and the corresponding calculated oscillations 16.9 C.

The annual air temperature oscillations at gauge TW2 were between −1.8 C and 32.1
◦ ◦
C, so that their amplitude was 33.9 C, whereas the annual water temperature oscillations
◦ ◦
were from 2.7 to 19.0 C and the amplitude was 16.3 C. However, the annual water

temperature oscillations at gauge TW1 were from 2.9 to 16.6 C, and the amplitude was 13.7
◦ ◦
C. At gauge TC5 the measured annual oscillations were 13.8 C, and the corresponding

calculated oscillations 16.8 C.
The annual air temperature oscillations close to gauge TC6 were, for the selected
◦ ◦
period, between −5.8 and 30.1 C, so that their amplitude was 35.9 C. The annual
concrete temperature oscillations at gauge TC6 were significantly smaller (measured
◦ ◦
24.7 C and calculated 27.5 C).
3.3. Temperature Field of the Dam
Based on the results of the 2D thermal analyses of the complete dam for the one year
period, the temperature fields of the dam during the year were defined, using time periods of
two months, starting from 1 October 2013 at 12:00. The temperature fields were plotted using
the computer program DIANA [49], where an identical model of the dam as in MATLAB was
constructed (i.e., it had the same arrangement and the same denotation of finite elements
and nodes), so that when making the temperature fields in DIANA, the results of the thermal
analyses calculations from MATLAB (the nodal temperatures at the selected times) were
used. Comparing the calculated temperature fields, it was found that the largest gradients in
the concrete temperature were recorded in the area close to the surface of the insolated
upper part of the downstream side of the dam (Figure 14). This was in summer, when the
highest air temperatures occurred and when not much shade was recorded on the
downstream side of the dam. For determining the temperature gradient by depth, the two
typical cross-sections, A–A and C–C, were chosen (Figures 2 and 14).
Appl. Sci. 2021, 11, x FOR PEER REVIEW 16 of 20 The four selected points (nodes) coincided with the locations of the temperature
gauges.

Figure 14.The calculated fieldof the Moste Dam (1 August 2014at 12:00) .
Figure 14 .Thecalculatedtemperaturerature field of theMoste Dam (1August 2014 at12:00) .

35
A - A Summer - Calculated A - A Winter - Calculated
30
A - A Summer - Measured A - A Winter - Measured
C - C Summer - Calculated C - C Winter - Calculated
Appl. Sci. 2021, 11, 705 15 of 19

The concrete temperatures at the two characteristic cross sections in summer (1 August
2014 at 12:00) and in winter (1 February 2014 at 12:00) are presented in Figure 15. It can be
seen from this figure that in cross-section A–A at the downstream insolated side of the dam in
the summer the calculated temperature from the concrete surface to a depth of 0.2 m was
◦ ◦
reduced by 9.8 C, whereas it was reduced at a depth of 1.0 m by 2.4 C, and by an additional

2.0 C at a depth of 2.0 m. This suggests that the temperature gradient is largest in a very
narrow area along the surface of the concrete. The calculated concrete temperatures in
cross-section A–A also indicate that the temperature stabilizes at a depth of about 6.0 m. In
the analyzed case, in winter the temperature gradient of the calculated temperatures close to
the surface was much smaller, but the strong effect of temperature can be observed up to a
depth of 2.0 m. Both in summer and in winter the measured values at three points (at depths

of 0.2, 0.5 and 1.0 m) were somewhat higher (by 1.3 to 2.5 C)
Figure 14. The calculated temperature field of the Moste Dam (1 August 2014 at 12:00).

than the calculated values.

35
A - A Summer - Calculated A - A Winter - Calculate
30
A - A Summer - Measured A - A Winter - Measured
C - C Summer - Calculated C - C Winter - Calculate
25
C - C Summer - Measured C - C Winter - Measured
20
(°C)

15
Temperature

10

-5
0 1 2 3 4 5
Depth (m)
Figure15. The concretetemperatures at the two typical cross-sections in summer and in winter.
15. The at the two - in .

In cross-section C–C at the upstream side in the summer, when the water level of 3.4. Uncertainty Analysis
the reservoir was above the analyzed cross-section, the calculated temperatures from the
As part of the thermal analyses, uncertainty analyses ◦ of the results of the calcula- concrete surface to a depth of
0.3 m were reduced by 2.4 C, whereas up to a depth of 1.4 m tions of the concrete temperatures were ◦also performed,
where six normally distributed they were reduced by an additional 1.7 C. This suggests that the temperature gradient
random variables were used to determine the dispersion of the results. Their expected is the greatest up to a depth of
0.3 m, and is significantly lower in comparison with the values and standard deviations are given in Table 1. Thermal
analysis was carried◦ out for analyzed case in winter (the temperature gradient up to a depth of 0.3 m was 5.8 C) when
a typical one-month period in the summer (from 1 July 2014 at 12:00 to 1 August 2014 at the water level of the reservoir was below
the analyzed cross-section, so that the concrete
12:00), where appropriate temperatures, obtained when calculating the temperature surface was exposed to the air. Both in
summer and in winter the measured value at a
dam
fields of the during a one-year period, were considered as the initial temperatures ◦ in depth of 0.3 matched the calculated values very well (the differences
were 0.1 C and

the analysis (Section 3.3). The concrete temperatures (at 12 selected points on and inside 0.6 C, respectively).
the dam; Figures 2 and 14, Table 5) were analyzed at the end of a period of one month. Sixty 3.4.UncertaintyrepetitionsAnalysisofthecalculation (for 60 different combinations of

values of six normally distributedAspartrandomofthethermalvariables),analyses,whichuncertaintyresultedinanalyses12differentofthesamples,resultsofwerethe calculationsperformed.

Theofthecalculatedconcrete valuestemperaturesofsevenwerestatisticsalsoperformed,ofselectedwheresamplessix normallyoftheconcretedistributedtemperaturesrandom

arevariablesshownwereinTableusd6,towhereasdetrminthe correspondingthedispersion boxofthe-


andresult-whiskers.Theirplotsexpectedarepresentedvalues inandFigurestandard16. deviations
are given in Table 1. Thermal analysis was carried out for a typical one-month period in the
summer (from 1 July 2014 at 12:00 to 1 August 2014 at 12:00), where appropriate temperatures,
obtained when calculating the temperature fields of the dam during a one-year period, were
considered as the initial temperatures in the analysis (Section 3.3). The concrete temperatures (at
12 selected points on and inside the dam; Figures 2 and 14, Table 5) were analyzed at the end of a
period of one month. Sixty repetitions of the calculation (for 60 different combinations of values of
six normally distributed random variables), which resulted in 12 different samples, were performed.
A1 A2 A3 A4 A5 B1 B2 B3 B4 B5 C1 C2
Depth (m) 0.0 0.2 0.5 1.0 2.0 0.0 1.6 3.2 4.8 6.4 0.0 0.3

The maximum average temperature of the 12 samples of populations was obtained


at the point A1 (30.24 °C), and the minimum at the point B5 (10.40 °C). The maximum
Appl. Sci. 2021, 11, 705 16 of 19
standard deviations were also obtained at point A1 (1.15 °C), whereas the minimum
standard deviations were identified at points B5, C1 and C2 (0.08 °C). It was found that
the average temperatures of 12 samples of the populations, as compared with the calcu-
The calculated values of seven statistics of selected samples of the concrete temperatures lated temperatures at the selected points, without considering
the point of view of un-

are shown in Table 6, whereas the corresponding box-and-whiskers plots are presented in
certainty, differed by 0.1 °C at four points (A1, A3, A4, B1), whereas at the other eight

Figure 16.
points there was no difference. The maximum temperature of 12 samples of the popula-

tions was obtained at the point A1 (32.6 °C) and the minimum at the point B5 (10.3 °C). At
Table 5. Locations of the 12 selected points at typical cross-sections.
point A1 the maximum interquartile range (1.83 °C) and the maximum range (4.5 °C)
were also registered. The minimum interquartile range was calculated for the point C1
Downstream–Above Downstream–Below Upstream

Point (0.03 °C), whereas the minimum range was recorded at the points B5, C1 and C2 (0.3 °C).
A1 A2 A3 A4 A5 B1 B2 B3 B4 B5 C1 C2
Depth (m) 0.0 0.2 0.5 1.0 2.0 0.0 1.6 3.2 4.8 6.4 0.0 0.3
Table 6. The calculated values of 7 statistics of the concrete temperatures at 12 selected points.

Temperature (°C)
Point
Table 6. The calc ulated values of *7 statistics of the concrete temperatures at 12 selected points.
4
n 1 X 2 σX 3 X min
4
T1 5 Q2
5
T3 5 X max
A1 60 30.24 1.15 28.1 Température29.20( C) 30.25 31.03 32.6

Point 1 * 3 4 5 5 5 4
A2
n 60 20.34 0.37 19.6 20.10 20.30 20.60 21.1

X σX X
min T
1 Q
2 T
3 X
max

A3 60 18.48 0.22 18.1 18.30 18.50 18.60 18.9


A1 60 30.24 1.15 28.1 29.20 30.25 31.03 32.6

A4 60 18.00 0.17 17.7 17.90 18.00 18.10 18.3


A2 60 20.34 0.37 19.6 20.10 20.30 20.60 21.1

A3 60 18.48 0.22 18. 18.30 18.50 18.60 18.9


A5 60 15.95 0.19 15.5 15.80 15.95 16.10 16.3

A4 60 18.00 0.17 17.7 17.90 18.00 18.10 18.3


B1 60 23.28 0.60 22.0 22.90 23.30 23.70 24.7

A5 60 15.95 0.19 15.5 15.80 15.95 16.10 16.3


B2 60 15.93 0.14 15.6 15.80 15.90 16.00 16.2

B1 60 23.28 0.60 22.0 22.90 23.30 23.70 24.7


B3 60 12.79 0.17 12.5 12.68 12.80 12.90 13.2
B2 60 15.93 0.14 15.6 15.80 15.90 16.00 16.2

B3 B4 60 60 1211.79.13 0.017.13 1210.5.9 1211.68. 00 12.1180.10 12.9011.20 13.211.5


B4 60 11.13 0.13 10.9 11.00 11. 11.20 11.5
B5 60 10.40 0.08 10.3 10.30 10.40 10.40 10.6

B5 60 10.40 0.08 10.3 10.30 10.40 10.40 10.6


C1 60 17.21 0.08 17.1 17.20 17.20 17.23 17.4

C1 60 17.21 0.08 17.1 17.20 17.20 17.23 17.4


C2 60 14.82 0.08 14.7 14.80 14.80 14.90 15.0
C2 60 14.82 0.08 14.7 14.80 14.80 14.90 15.0

*
n: sample size. sample:samplemean. meanσ:standard. deviatio n:standardof deviation thesample.of the sample.,:minimumX,X andmax : minimummaximum valu eandmaximumofthesamplevalue

1 1 2 2 3 3 4 4

n: sample size. X : X σ X X


max

X X min
elements. 5 T1, T2 , T3 : first, second5 and third quartile of the sample elements.
of the sample elements. Q1, Q2, Q3: first, second and third quartile of the sample elements.
35

30

25

20

15

10
A1A2 A3 A4 A5 B1 B2 B3 B4 B5
C1 C2
Location

FigureFigure16. Box16.Box-and-and-whiskers-whiskersplotsplotsftheof theconconcretetemperatures12atselected12selectedpointspoints..

The maximum average temperature of the 12 samples of populations was obtained at


◦ ◦
the point A1 (30.24 C), and the minimum at the point B5 (10.40 C). The maximum standard

deviations were also obtained at point A1 (1.15 C), whereas the minimum standard

deviations were identified at points B5, C1 and C2 (0.08 C). It was found that the average
temperatures of 12 samples of the populations, as compared with the calculated
temperatures at the selected points, without considering the point of view of uncertainty,

differed by 0.1 C at four points (A1, A3, A4, B1), whereas at the other eight points there was
no difference. The maximum temperature of 12 samples of the populations was obtained at
◦ ◦
the point A1 (32.6 C) and the minimum at the point B5 (10.3 C). At point
Appl. Sci. 2021, 11, 705 17 of 19

◦ ◦
A1 the maximum interquartile range (1.83 C) and the maximum range (4.5 C) were also

registered. The minimum interquartile range was calculated for the point C1 (0.03 C),

whereas the minimum range was recorded at the points B5, C1 and C2 (0.3 C).
4. Conclusions
This article presents a relatively simple procedure for modeling the heat transfer
process in a concrete dam, taking into account time-varying boundary conditions on the
upstream and downstream sides of the dam (i.e., the water level of the reservoir,
spillover, insolation, and shading) which affect the dam’s thermal fields.
The validity of the procedure was verified on the example of the large arch–gravity
concrete Moste Dam (the highest such dam in Slovenia), where a sophisticated
automated system has recently been installed. This system is used for the measurement
of concrete temperatures, as well as to monitor the surrounding conditions (i.e., the water
temperatures, the water level of the reservoir, the height of spillover, the air temperature
and the solar insolation).
Thermal analyses (1D and 2D) for the analysis of non-linear and non-stationary heat
conduction through solids were performed using a FEM-based program, which was
complemented by two specially developed programs for determining the effects of
boundary conditions during the analyzed 15-day period, as well as during a period of one
year. Uncertainty analyses were also carried out to estimate inherent random effect.
It was found that the results of the performed thermal analyses fitted in well with the
experimentally determined concrete temperature measurements. The results showed
that at the insolated side of the dam, the temperature gradient was largest in a very
narrow area along the concrete surface, but the temperature did not stabilize at depths
shallower than about 6 m.
The obtained results (i.e., seasonal temperature distributions in the dam) affect the
thermal loads, which can cause high tensile stresses in concrete and importantly affect
the occurrence and behavior of cracks, as well as displacements; especially in arch
dams and arch-gravity dams.

Author Contributions: Conceptualization, P.Ž., G.T. and A.K.; méthodologie, P.Ž., G.T. et A.K.;
software, P.Ž. and G.T.; validation, P.Ž. and G.T.; formal analysis, P.Ž.; enquête, P.Ž., G.T. et
A.K.; resources, P.Ž.; data curation, P.Ž.; writing—original draft preparation, P.Ž., G.T. and A.K.;
writing—review and editing, P.Ž., G.T. and A.K.; visualization, P.Ž. and G.T.; supervision, P.Ž.,
G.T. et A.K.; project administration, P.Ž.; funding acquisition, P.Ž. All authors have read and
agreed to the published version of the manuscript.
Funding: The presented article is part of research work that was performed within the scope of
the first author’s doctoral studies, and was partly funded by the European Union through the
European Social Found.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author. The data are not publicly available due to the requirements of the company
Savske elektrarne Ljubljana Ltd.
Acknowledgments: The authors would like to thank the company Savske elektrarne Ljubljana
Ltd., which enabled the field work. The authors would also like to acknowledge the financial
support of the Slovenian Research Agency through infrastructure program I0-0032, research
program P2-0260 and research project J2-9196.
Conflicts of Interest: The authors declare no conflict of interest.

References

1. Bukenya, P.; Moyo, P.; Beushausen, H.; Oosthuizen, C. Health monitoring of concrete dams: A literature review. J. Civ. Struct.
Health Monit. 2014, 4, 235–244. [CrossRef]
Appl. Sci. 2021, 11, 705 18 of 19

2. Santillan, D.; Salete, E.; Toledo, M.A. A methodology for the assessment of the effect of climate change on the thermal-strain-
stress behaviour of structures. Eng. Struct. 2015, 92, 123–141. [CrossRef]
3. Dilger, W.H.; Ghali, A.; Chan, M.; Cheung, M.S.; Maes, M.A. Temperature stresses in composite box girder bridges. J. Struct.
Eng. 1983, 109, 1460–1478. [CrossRef]
4. Carslaw, H.S.; Jaeger, J.C. Conduction of Heat in Solids, 2nd ed.; Oxford University Press: New York, NY, USA, 1986; ISBN
978-0198533689.
5. Leger, P.; Venturelli, J.; Bhattacharjee, S.S. Seasonal temperature and stress distributions in concrete gravity dams. Part 1:
Modelling. Can. J. Civ. Eng. 1993, 20, 999–1017. [CrossRef]
6. Leger, P.; Venturelli, J.; Bhattacharjee, S.S. Seasonal temperature and stress distributions in concrete gravity dams. Part 2:
Behavior. Can. J. Civ. Eng. 1993, 20, 1018–1029. [CrossRef]
7. U.S. Army Corps of Engineers. Engineering and Design: Arch Dam Design (Engineer Manual EM 1110-2-2201); Military
Bookshop: Washington, DC, USA, 1994; ISBN 978-1780397610.
8. U.S. Army Corps of Engineers. Engineering and Design: Gravity dam design (Engineer manual EM 1110-2-2200); Military
Bookshop: Washington, DC, USA, 1995.
9. International Commission on Large Dams. Automated Dam Monitoring Systems: Guidelines and Case Histories—Bulletin 118;
Commission Internationale des Grands Barrages: Paris, France, 2000.
10. International Commission on Large Dams. Dam Surveillance Guide—Bulletin 158; Commission Internationale des Grands
Barrages: Paris, France, 2012.
11. Saetta, A.; Scotta, R.; Vitaliani, R. Stress analysis of concrete structures subjected to variable thermal loads. J. Struct. Eng.
1995, 121, 446–457. [CrossRef]
12. Daoud, M.; Galanis, N.; Ballivy, G. Calculation of the periodic temperature field in a concrete dam. Can. J. Civ. Eng. 1997, 24,
772–784. [CrossRef]
13. Wu, Y.; Luna, R. Numerical implementation of temperature and creep in mass concrete. Finite Elem. Anal. Des. 2001, 37, 97–
106. [CrossRef]
14. Noorzaei, J.; Bayagoob, K.H.; Thanoon, W.A.; Jaafar, M.S. Thermal and stress analysis of Kinta RCC dam. Eng. Struct. 2006,
28, 1795–1802. [CrossRef]
15. Sheibany, F.; Ghaemian, M. Effects of environmental action on thermal stress analysis of Karaj concrete arch dam. J. Eng.
Mech. 2006, 132, 532–544. [CrossRef]
16. Agullo, L.; Mirambell, E.; Aguado, A. A model for the analysis of concrete dams due to environmental thermal effects. Int. J.
Numer. Methods Heat Fluid Flow 1996, 6, 25–36. [CrossRef]
17. Leger, P.; Leclerc, M. Hydrostatic, temperature, time-displacement model for concrete dams. J. Eng. Mech. 2007, 133, 267–
277. [CrossRef]
18. Li, F.Q.; Wang, Z.Y.; Liu, G.H.; Fu, C.J.; Wang, J.J. Hydrostatic seasonal state model for monitoring data analysis of concrete
dams. Struct. Infrastruct. Eng. 2015, 11, 1616–1631. [CrossRef]
19. Zhang, P.X.; Ayari, M.L.; Robinson, L.C. Role of fundamental heat transfer frequency in the computation of transient thermal
fields in massive structures. Can. J. Civ. Eng. 1997, 24, 1059–1065. [CrossRef]
20. Leger, P.; Cote, M.; Tinawi, R. Thermal protection of concrete dams subjected to freeze—Thaw cycles. Can. J. Civ. Eng.
1995, 22, 588–602. [CrossRef]
21. Ardito, R.; Maier, G.; Massalongo, G. Diagnostic analysis of concrete dams based on seasonal hydrostatic loading. Eng.
Struct. 2008, 30, 3176–3185. [CrossRef]
22. Leger, P.; Seydou, S. Seasonal thermal displacements of gravity dams located in northern regions. J. Perform. Constr. Facil.
2009, 23, 166–174. [CrossRef]
23. Xu, B.S.; Liu, B.B.; Zheng, D.J.; Chen, L.; Wu, C.C. Analysis method of thermal dam deformation. Sci. China Technol. Sci.
2012, 55, 1765–1772. [CrossRef]
24. Mata, J.; de Castro, A.T.; da Costa, J.S. Time-frequency analysis for concrete dam safety control: Correlation between the
daily variation of structural response and air temperature. Eng. Struct. 2013, 48, 658–665. [CrossRef]
25. Malm, R.; Ansell, A. Cracking of concrete buttress dam due to seasonal temperature variation. ACI Struct. J. 2011, 108, 13–22.
26. Maken, D.D.; Leger, P.; Roth, S.N. Seasonal thermal cracking of concrete dams in northern regions. J. Perform. Constr. Facil.
2014, 28. [CrossRef]
27. Labibzadeh, M.; Sadrnejad, S.A.; Khajehdezfuly, A. Thermal assessment of Karun-1 Dam. Trends Appl. Sci. Res. 2010, 5,
251–266. [CrossRef]
28. Santillan, D.; Salete, E.; Toledo, M.A. A new 1D analytical model for computing the thermal field of concrete dams due to the
environmental actions. Appl. Therm. Eng. 2015, 85, 160–171. [CrossRef]
29. Jin, F.; Chen, Z.; Wang, J.T.; Yang, J. Practical procedure for predicting non-uniform temperature on the exposed face of arch
dams. Appl. Therm. Eng. 2010, 30, 2146–2156. [CrossRef]
30. Santillan, D.; Salete, E.; Vicente, D.J.; Toledo, M.A. Treatment of solar radiation by spatial and temporal discretization for
modeling the thermal response of arch dams. J. Eng. Mech. 2014, 140, 05014001. [CrossRef]
31. Mirzabozorg, H.; Hariri-Ardebili, M.A.; Shirkhan, M.; Seyed-Kolbadi, S.M. Mathematical modeling and numerical analysis of
thermal distribution in arch dams considering solar radiation effect. Sci. World J. 2014, 597393. [CrossRef]
Appl. Sci. 2021, 11, 705 19 of 19

32. Liu, H.B.; Liao, X.W.; Chen, Z.H.; Zhang, Q. Thermal behavior of spatial structures under solar irradiation. Appl. Therm. Eng.
2015, 87, 328–335. [CrossRef]
33. An, G.; Yang, N.; Li, Q.B.; Hu, Y.; Yang, H.T. A simplified method for real-time prediction of temperature in mass concrete at
early age. Appl. Sci. 2020, 10, 4451. [CrossRef]
34. Hu, Y.; Zuo, Z.; Li, Q.B.; Duan, Y.L. Boolean-based surface procedure for the external heat transfer analysis of dams during
construction. Math. Probl. Eng. 2013, 175616. [CrossRef]
35. Liu, Y.; Zhang, G.X.; Zhu, B.F.; Shang, F. Actual working performance assessment of super-high arch dams. J. Perform.
Constr. Facil. 2016, 30, 04015011. [CrossRef]
36. Yin, T.; Li, Q.B.; Hu, Y.; Yu, S.; Liang, G.H. Coupled thermo-hydro-mechanical analysis of valley narrowing deformation of
high arch dam: A case study of the Xiluodu project in China. Appl. Sci. 2020, 10, 524. [CrossRef]
37. Xia, Y.; Chen, B.; Zhou, X.Q.; Xu, Y.L. Field monitoring and numerical analysis of Tsing Ma Suspension Bridge temperature
behavior. Struct. Control. Health Monit. 2013, 20, 560–575. [CrossRef]
38. Su, J.Z.; Xia, Y.; Ni, Y.Q.; Zhou, L.R.; Su, C. Field monitoring and numerical simulation of the thermal actions of a supertall
structure. Struct. Control. Health Monit. 2017, 24, e1900. [CrossRef]
39. Abid, S.R.; Mussa, F.; Taysi, N.; Ozakca, M. Experimental and finite element investigation of temperature distributions in
concrete-encased steel girders. Struct. Control. Health Monit. 2018, 25, e2042. [CrossRef]
40. Li, B.; Yang, J.; Hu, D.X. Dam monitoring data analysis methods: A literature review. Struct. Control. Health Monit. 2020, 27,
e2501. [CrossRef]
41. Žvanut, P.; Turk, G.; Kryžanowski, A. Effects of changing surrounding conditions on the thermal analysis of the Moste
concrete dam. J. Perform. Constr. Facil. 2016, 30, 04015029. [CrossRef]
42. Zienkiewicz, O.C. The Finite Element Method, 3rd ed.; McGraw-Hill: London, UK, 1977; ISBN 978-0070840720.
43. Bofang, Z. Prediction of water temperature in deep reservoirs. Dam Eng. 1997, 8, 13–25.
44. Wolfram Mathematica Home Page. Available online: http://www.wolfram.com/mathematica (accessed on 10 August 2020).
45. MathWorks MATLAB Home Page. Available online: http://www.mathworks.com/products/matlab (accessed on 11 July 2020).
46. Žvanut, P.; Vezoˇcnik, R.; Turk, G.; Ambrožiˇc, T. Determination of the shading of the downstream surface of the Moste
concrete dam. Geod. Vestnik 2014, 58, 453–465. [CrossRef]
47. Ilc, A. Nonlinear Analysis of Massive Concrete at Successive Construction. Ph.D. Thesis, University of Ljubljana, Faculty of
Civil and Geodetic Eng., Ljubljana, Slovenia, 2013.
48. Žvanut, P. Thermal Analysis of Large Arch-Gravity Concrete Dams. Ph.D. Thesis, University of Ljubljana, Faculty of Civil and
Geodetic Eng., Ljubljana, Slovenia, 2017.
49. DIANA FEA Home Page. Available online: https://dianafea.com (accessed on 11 August 2020).

Vous aimerez peut-être aussi